Download A pathway toward safer anesthesia: Stereochemical advances

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts
no text concepts found
Transcript
AANA Journal Course
Update for Nurse Anesthetists
6
*6 CE Credits
A pathway toward safer anesthesia:
Stereochemical advances
Joseph A. Joyce, CRNA, BS
Greensboro, North Carolina
Advances in organic analytical chemistry recently have led
to the ability to produce medications that are stereochemically pure. At present, most medications commercially available are racemic mixtures, which are 50%-50% mixtures of
the constituent stereoisomers. The human body is a chiral
system and, as a result, exhibits stereoselectivity in its utilization of medications. When a racemic mixture is administered, all the stereoisomers must be metabolized. Development of stereochemically pure medications can lead to safer
medications, reduce the amount of medication needed to
produce the desired effect, and, potentially, reduce untoward and/or toxic effects of a medication.
Key words: Chiral, racemic mixture, enantiomers, geometric
isomers, optical activity, stereoselectivity.
OBJECTIVES
At the completion of this course, the reader should be
able to:
1. List the types of stereoisomers.
2. Differentiate among stereoisomers.
3. Identify the implications of using racemic mixtures as opposed to stereochemically pure medications.
4. List 2 potential benefits of stereochemically pure
medications.
5. List 2 stereochemically pure medications.
Almost every time an anesthetist administers a medication, 2 or more medications are, in fact, given. That
would seem to be quite an inflammatory statement,
one that requires an explanation. How can that possibly be true? Only one syringe or vaporizer is used, and
those are clearly and correctly labeled as to the contents. How, then, could 2 or more medications have
been administered? Can such a statement be accurate?
The simple answer is: Yes! The purpose of this article is
to present an overview of a “new” avenue for pharmacological research and development. Such an overview
must, of necessity, begin with a review of chemistry.
Chemistry is divided into 2 broad headings: the first
is inorganic chemistry and the second, organic chemistry. Organic chemistry is concerned primarily with
the composition, formation, and reactions of carbonbased compounds produced within or as a by-product
of either the plant or the animal kingdoms. It is the area
of organic chemistry with which the anesthetist is most
concerned and from which the answers to the initial
questions will be gleaned.
Within the broad area of organic chemistry are
numerous, more specialized subdivisions, such as
stereochemistry, which is the focus of this article. Stereochemistry is the branch of organic chemistry that
deals with the spatial arrangement of atoms’ and molecules’ attachment at a specific carbon atom and the
resultant relationship to the physical properties exhibited by the molecule as a whole. Table 11 is a compilation of definitions of terms specific to stereochemistry.
In any basic study of chemistry, one learns the concept of isomers. Isomers are chemical compounds that
have identical empirical formulas but are structurally
different. In anesthesia, we probably are most familiar
with this concept through the volatile anesthetics,
enflurane and isoflurane. Both of these anesthetics have
the same empirical formula, C3H2ClF5O, but the struc-
* The American Association of Nurse Anesthetists is accredited as a provider of continuing education in nursing by the American Nurses Credentialing
Center Commission on Accreditation. The AANA Journal course will consist of 6 successive articles, each with objectives for the reader and sources for
additional reading. At the conclusion of the 6-part series, a final examination will be printed in the AANA Journal. Successful completion will yield the
participant 6 CE credits (6 contact hours), code number: 23633, expiration date: July 31, 2002.
AANA Journal/February 2002/Vol. 70, No. 1
63
Table 1. Stereochemistry terms and definitions1
Term
Definition
Achiral
A molecule devoid of chirality
Antipode
The opposing optical isomer
Chiral (asymmetrical) center
Typically refers to a carbon atom to which 4 different atomic or molecular substituents are bonded
Diastereoisomer
(diastereomer)
Stereoisomers with 2 chiral centers physiochemically different from one another
Distomer
Enantiomer with the least activity
Enantiomer
One of the mirror images
Eutomer
Enantiomer with the majority of a given effect
Geometric isomers
Molecules formed by limitation of the rotation of molecules and/or atoms around a carbon atom;
restriction by either a double bond or a rigid carbon ring system
L- and – and
D- and +
Descriptors of optical activity: L- and – refer to levorotary enantiomers; D- and + refer to dextrrotary enantiomers
Optical activity
A molecule that can change the direction of plane-polarized light
Racemic mixture
An equal proportion of stereoisomers; a solution containing the possible stereoisomers in
equilibrium
S and R
Absolute configuration or spatial arrangement of optical isomers independent of the direction of
optical activity
Stereoselective
Preferentially related to one stereoisomer
Stereospecific
Specific to only one stereoisomer
Tautomeres
Conformational isomeric forms that exist in equilibrium but change from one form to another with
ease depending on the physical conditions surrounding them
tural placement of the chlorine atom (Cl) yields entirely
different compounds. Figure 1 shows the 2-dimensional, structural formulas for enflurane and isoflurane.
There are 3 types of stereoisomers: optical, geometric, and conformational. Optical isomers, or enantiomers, are compounds that have identical empirical
and structural formulas; however, enantiomers differ in
the 3-dimensional spatial arrangement of the atoms
and/or molecules bonded to the carbon atom. Actually,
these forms of stereoisomers are mirror images of each
other that cannot be superimposed on one another, just
as one’s left hand cannot be superimposed on the right
hand. Figures 2 and 3 demonstrate the spatial differentiation of optical isomers.
Not all organic compounds are capable of forming
optical isomers. For an organic molecule to have optical
isomers, that molecule must have at least 1 chiral carbon,
sometimes called an asymmetrical carbon. The word chiral comes from the Greek word, chiros,2 which translates
as “hand;” therefore, a chiral compound may be righthanded or left-handed. Whether a molecule is right- or
left-handed is determined by the relative strength of
forces exerted on the chiral carbon by the atoms and/or
molecules bonded to it. A right-handed molecule is des64
AANA Journal/February 2002/Vol. 70, No. 1
Figure 1. Structural isomers
Empirical chemical formula: C3H2CIF5O (for both
enflurane and isoflurane)
Structural formulas
Isoflurane
Enflurane
ignated by the letter “R” for rectus; a left-handed molecule is designated by the letter “S” for sinister.
In addition to spatial or 3-dimensional differences,
optical isomers differ as to the direction in which they
cause plane-polarized light to rotate. One enantiomer will
produce counterclockwise rotation; the other will produce clockwise rotation of this type of light. The direction
to which the light is rotated is symbolically designated as
follows: counterclockwise rotation is termed levorotatory
Figure 2. Generic optical isomers
Figure 4. Geometric isomers
cis- or Z
Figure 3. Isoflurane optical isomers
trans- or E
Figure 5. Conformational isomers
“Beach-chair” conformation
and symbolized by an “L-” or “–” notation; clockwise
rotation of plane-polarized light is termed dextrorotatory
and is symbolized by a “D-” or “+” notation. Thus, enantiomers have 2 independent descriptive designations.
There is no relationship between the handedness of a
molecule and the direction to which the molecule causes
plane-polarized light to rotate.3,4
Geometric isomers constitute the second group of
stereoisomers. This type of stereoisomer arises when
molecular rotation around a carbon atom is restricted
or prohibited by a carbon-to-carbon double bond or by
the presence of a rigid carbon ring system. With single
carbon-to-carbon bonds, the attached molecules or
atoms are able to rotate rather freely around the carbon
atom. Carbon-to-carbon double bonds are shorter and
more rigid than single bonds. Visually, the rotation
around a carbon single bond is similar to the rotational
ability of the humerus, whereas the rotation around a
carbon-to-carbon double bond is similar to the rotational ability of the radius and ulna, which is somewhat
more restricted, more rigid. Geometric isomers have
been differentiated, traditionally, by the following prefixes: cis-, from Greek, which translates as “same,” and
trans-, also from Greek, which translates as “opposite.”
More recently, these designations have been indicated
by the letters “Z-” and “E-” for the German words,
zusamman and entgegen, respectively. The Z- and Eroughly correspond, respectively, to the traditional cisand trans- nomenclature, although both methods con-
“Boat” conformation
tinue to be used. Figure 4 shows a generic representation of geometric isomers.
Conformational isomers, or conformers, are the last
subset of stereoisomers. These isomers have nonidentical spatial arrangements of the atoms and/or molecules
bonded to a particular carbon. This subset of stereoisomers occurs as a result of the rotation of the bonded
constituents around 1 single carbon bond or around
more than 1 single carbon bond. Essentially, this molecule must twist, contort, or fold to bind to a target
receptor and produce the desired effect. Figure 5 is a
representation of conformational isomers.
Biological significance
Biological systems are, of a sort, living chemical laboAANA Journal/February 2002/Vol. 70, No. 1
65
ratories, some obviously more complex than others.
Arguably the most complex such living laboratory is
the human body. In most plants and animals, enzymes
and other biologically active molecules contain a chiral
carbon, sometimes more than one such carbon. The
human body can be considered a collection of chiral
compounds and, therefore, a chiral system.5 For example, to be used in protein synthesis, amino acids must
be L- enantiomers.6 Receptor sites are protein molecules built from specific enantiomers and are optically
active. Thus, receptors are able to distinguish between
stereoisomers, that is, they exhibit stereoselectivity, so
that only medications with the “proper” optical activity,
geometric structure, or conformation are able to interact with the receptor protein.
Traditionally, medications developed in a laboratory,
whether chiral, geometric, or conformational compounds, have been racemic mixtures, which are solutions containing all the possible isomers in equilibrium.
Therefore, a single vial or ampule of such a medication
may contain 2 or more stereoisomers or medications.
Biologically active molecules within the body are able to
distinguish between stereoisomers.6,7 As a result of the
body’s ability to distinguish between stereoisomers,
administration of a racemic mixture can be viewed as
administering 2 or more medications, all of which may
behave differently with regard to toxicology, pharmacodynamics, and pharmacokinetics (Table 2).1,8
During the 1990s, advances in analytical chemistry
specifically applied to medication design have allowed
scientists to more easily separate and manufacture
specifically oriented molecules. In other words, rather
than producing racemic mixtures, a stereochemically
pure compound can be produced. Recognizing the
advances in analytical chemistry technology in recent
years, the US Food and Drug Administration, in 1992,
issued guidelines7 for pharmaceutical companies
regarding investigation of the pharmacological properties of various medications. In essence, these guidelines
put forth 3 cases for specific testing of pure stereoisomers:7 (1) both stereoisomers demonstrate desirable
effects, (2) one stereoisomer demonstrates pharmacological activity, while the other does not, and (3) the
stereoisomers demonstrate completely different pharmacological activities or have a different concentrationresponse relationship for a certain property.
Anesthesia implications
One of the goals of studying, developing, and marketing a medication composed of a single, pure stereoisomer is to produce a decided benefit for the patient who
receives that medication. For many medications, one
stereoisomer will exhibit greater potency than another.
Greater potency often translates into a reduction in the
66
AANA Journal/February 2002/Vol. 70, No. 1
Table 2. Some achiral and chiral anesthesia-related
drugs8
Achiral
Chloroprocaine
Neostigmine
Dopamine
Nitrous oxide
Edrophonium
Propofol
Fentanyl
Sevoflurane
Gallamine triethiodide
Tetracaine
Lidocaine
Chiral
Alcuronium
Hyoscine
Atropine
Isoflurane
Bupivacaine
Ketamine
Cisatracurium
Levobupivacaine
Desflurane
Mepivacaine
Dexmedetomidine
Methohexital
Dobutamine
Morphine
Enflurane
Prilocaine
Etidocaine
Remifentanil
Etomidate
Ropivacaine
Fenoldopam
Thiopental
Glycopyrrolate
Tubocurarine
Halothane
amount required to produce the desired effect. Often,
one stereoisomer will demonstrate a greater safety
index compared with the racemate(s) or with the
racemic mixture. For the anesthetist, both of these are
highly sought-after potential patient benefits. Increasing the safety index of an anesthesia medication is a
benefit to all involved, patients and practitioners. There
are several examples of newly developed medications
in which one stereoisomer demonstrates greater pharmacological activity than another. One is dexmedetomidine. Dexmedetomidine is the dextrorotatory optical
isomer of the imidazole compound, medetomidine,
that acts specifically and selectively as an alpha2adrenoceptor agonist.9 Dexmedetomidine has demonstrated 8 times the specificity for alpha2-adrenoceptors
compared with clonidine.10 As a potent alpha2-adrenoceptor agonist, dexmedetomidine produces significant
sedation, analgesia, and anxiolysis while maintaining
hemodynamic stability without respiratory depression,
ease of arousal, and patient cooperation. Currently,
dexmedetomidine is marketed in the United States primarily for continuous infusion for intensive care
patients who require mechanical ventilation. Dexmedetomidine also significantly reduces anesthetic requirements,11 postoperative analgesic requirements,12 and
the vasoconstriction and shivering thresholds.13
An example of a pure stereoisomer that demonstrates
a greater safety index is levobupivacaine, or L-bupivacaine. Levobupivacaine is the most recently introduced
amide local anesthetic. Pharmacologically, L-bupivacaine seems to be equally potent with racemic bupivacaine, or bupivacaine, with onset of action and duration
of effect also similar to those of the racemic mixture.
The major issue regarding L-bupivacaine is the significantly greater safety index it has demonstrated thus far
and its reduced risk of toxic effects. For example, the
lethal dose of L-bupivacaine, from animal studies, has
been consistently 1.3 to 1.6 times greater than that for
the racemic mixture.14 Central nervous system toxic
effects seem to be less common with L-bupivacaine, and
larger doses are necessary to produce seizure activity or
apnea compared with racemic bupivacaine.14,15
Pure stereoisomeric medications can be an improvement particularly if metabolic products of a parent compound are toxic or deleterious, an example of which is
atracurium. Atracurium, introduced to anesthesia in
1982, is characterized by significant histamine release
when injected rapidly. Atracurium is eliminated by both
ester hydrolysis and Hofmann degradation. One
metabolite of atracurium is laudanosine, which is a
known epileptogenic compound. Higher plasma levels
of laudanosine tend to be formed when atracurium is
used via continuous infusion.16 Cisatracurium, or 1Rcis, 1'R-cis atracurium, is 1 of 10 geometric isomers of
atracurium. Cisatracurium is about 5 times more potent
than atracurium and produces significantly less histamine release than the racemic mixture. The predominant metabolic route of elimination of cisatracurium is
Hofmann degradation, with lesser elimination via ester
hydrolysis; however, cisatracurium results in 5 times
less laudanosine production than atracurium.17 This
represents a significant development when continuous
infusions are used, for example, with mechanically ventilated intensive care patients or patients undergoing
prolonged surgical procedures.
In addition to potentially increased drug potency and
safety indices, the development of stereochemically pure
medications may yield other benefits. The additional
benefits include simplification of dose-response relationships, reduced intersubject variability, and minimization
of toxic effects resulting from metabolism of the pharmacologically inactive or less active stereoisomer.18
of medications: stereochemically pure medications.
Many of these single isomer medications demonstrate
significantly greater potency and higher safety indices
compared with the complementary stereoisomers or
with the racemic mixtures from which they are derived.
By producing stereochemically pure medications,
potentially toxic metabolic products can be dramatically reduced, if not eliminated. The increased potency
may translate into a reduced dosage and reduce the
overall cost to the patient. Because of the high number
of anesthesia and anesthesia-related medications that
form stereoisomers, the anesthetist can look forward to
exciting advances in the safety of the medications that
are used on virtually a daily basis.
REFERENCES
1. Brocks DR, Jamali F. Stereochemical aspects of pharmacotherapy.
Pharmacotherapy. 1995;15:551-564.
2. Hutt AJ, Tan SC. Drug chirality and its clinical significance. Drugs.
1996;52(suppl 5):1-12.
3. Williams K, Lee E. Importance of drug enantiomers in clinical pharmacology. Drugs. 1985;30:333-354.
4. Karim A. Enantioselective assays in comparative bioavailability
studies of racemic drug formulations: nice to know or need to
know? J Clin Pharmacol. 1996;36:490-499.
5. Tracy TS. Stereochemistry in pharmacotherapy: when mirror images
are not identical. Ann Pharmacother. 1995;29:161-165.
6. Stryer L. Biochemistry. 3rd ed. New York, NY: W.H. Freeman.
1988:17.
7. FDA’s policy statement for the development of new stereoisomeric
drugs. Chirality. 1992;4:338-340.
8. Calvey TN. Isomerism and anaesthetic drugs. Acta Anaesthesiol
Scand. 1995;39(suppl 106):83-90.
9. Savola J-M, Virtanen R. Central alpha2-adrenoceptors are highly
stereoselective for dexmedetomidine, the dextro enantiomer of
medetomidine. Eur J Pharmacol. 1991;195:193-199.
10. Coughlan MG, Lee JG, Bosnjak ZJ Schmeling WT, Kampine JP,
Waltier DC. Direct coronary and cerebral vascular responses to
dexmedetomidine: Significance of endogenous nitric oxide synthesis. Anesthesiology. 1992;77:998-1006.
11. Aho M, Erkola O, Kallio A, Scheinin H, Korttila K. Dexmedetomidine infusion for maintenance of anesthesia in patients undergoing
abdominal hysterectomy. Anesth Analg. 1992;75:940-946.
12. Bhana N, Goa KL, McClellan KJ. Dexmedetomidine. Drugs.
2000;59:263-268.
13. Talke P, Tayefeh F, Sessler DI, Jeffrey R, Noursalehi M, Richardson C.
Dexmedetomidine does not alter the sweating threshold, but comparably and linearly decreases the vasoconstriction and shivering
thresholds. Anesthesiology. 1997;87:835-841.
14. Foster RH, Markham A. Levobupivacaine: a review of its pharmacology and use as a local anaesthetic. Drugs. 2000;59:551-579.
15. Thomas JM, Schug SA. Recent advances in the pharmacokinetics of
local anaesthetics: Long-acting amide enantiomers and continuous
infusions. Clin Pharmacokinet. 1999;36:67-83.
16. Chapple DJ, Miller AA, Ward JB, Wheatley PL. Cardiovascular and
neurological effects of laudanosine: Studies in mice and rats, and in
conscious and anaesthetized dogs. Br J Anaesth. 1987;59:218-225.
17. Smith CE, van Miert MM, Parker CJ, Hunter JM. A comparison of the
infusion pharmacokinetics and pharmacodynamics of cisatracurium,
the 1R-cis 1'R-cis isomer of atracurium, with atracurium besylate in
healthy patients. Anaesthesia. 1997;52:833-841.
18. Caldwell J. Importance of stereospecific bioanalytical monitoring in
drug development. J Chromatogr A. 1996;719:3-13.
Conclusion
The dramatic advances in analytical chemistry pertaining to stereochemistry have given rise to a new “class”
AUTHOR
Joseph A. Joyce, CRNA, BS, is a staff nurse anesthetist at Wesley Long
Community Hospital in Greensboro, NC.
AANA Journal/February 2002/Vol. 70, No. 1
67